23. Group Actions and Automorphisms Recall the Definition of an Action

Total Page:16

File Type:pdf, Size:1020Kb

23. Group Actions and Automorphisms Recall the Definition of an Action 23. Group actions and automorphisms Recall the definition of an action: Definition 23.1. Let G be a group and let S be a set. An action of G on S is a function G × S −! S denoted by (g; s) −! g · s; such that e · s = s and (gh) · s = g · (h · s) In fact, an action of G on a set S is equivalent to a group homomor- phism (invariably called a representation) ρ: G −! A(S): Given an action G × S −! S, define a group homomorphism ρ: G −! A(S) by the rule ρ(g) = σ : S −! S; where σ(s) = g · s. Vice-versa, given a representation (that is, a group homomorphism) ρ: G −! A(S); define an action G · S −! S by the rule g · s = ρ(g)(s): It is left as an exercise for the reader to check all of the details. The only sensible way to understand any group is let it act on some- thing. Definition-Lemma 23.2. Suppose the group G acts on the set S. Define an equivalence relation ∼ on S by the rule s ∼ t if and only if g · s = t for some g 2 G. The equivalence classes of this action are called orbits. The action is said to be transitive if there is only one orbit (neces- sarily the whole of S). Proof. Given s 2 S note that e · s = s, so that s ∼ s and ∼ is reflexive. If s and t 2 S and s ∼ t then we may find g 2 G such that t = g · s. But then s = g−1 · t so that t ∼ s and ∼ is symmetric. If r, s and t 2 S and r ∼ s, s ∼ t then we may find g and h 2 G such that s = g · r and t = h · s. In this case t = h · s = h · (g · r) = (hg) · r; so that t ∼ r and ∼ is transitive. 1 Definition-Lemma 23.3. Suppose the group G acts on the set S. Given s 2 S the subset H = f g 2 G j g · s = s g; is called the stabiliser of s 2 S. H is a subgroup of G. Proof. H is non-empty as it contains the identity. Suppose that g and h 2 H. Then (gh) · s = g · (h · s) = g · s = s: Thus gh 2 H, H is closed under multiplication and so H is a subgroup of G. Example 23.4. Let G be a group and let H be a subgroup. Let S be the set of all left cosets of H in G. Define an action of G on S, G × S −! S as follows. Given gH 2 S and g0 2 G, set g0 · (gH) = (g0g)H: It is easy to check that this action is well-defined. Clearly there is only one orbit and the stabiliser of the trivial left coset H is H itself. Lemma 23.5. Let G be a group acting transitively on a set S and let H be the stabiliser of a point s 2 S. Let L be the set of left cosets of H in G. Then there is an isomorphism of actions (where isomorphism is defined in the obvious way) of G acting on S and G acting on L, as in (23.4). In particular jGj jSj = : jHj Proof. Define a map f : L −! S by sending the left coset gH to the element g ·s. We first have to check that f is well-defined. Suppose that gH = g0H. Then g0 = gh, for some h 2 H. But then g0 · s = (gh) · s = g · (h · s) = g · s: Thus f is indeed well-defined. f is clearly surjective as the action of G is transitive. Suppose that f(gH) = f(gH). Then gS = g0s. In this case h = g−1g0 stabilises s, so that g−1g0 2 H. But then g and g0 are 2 in the same left coset and gH = g0H. Thus f is injective as well as surjective, and the result follows. Given a group G and an element g 2 G recall the centraliser of g in G is Cg = f h 2 G j hg = gh g: The centre of G is then Z(G) = f h 2 H j gh = hg g; the set of elements which commute with everything; the centre is the intersection of the centralisers. Lemma 23.6 (The class equation). Let G be a group. The cardinality of the conjugacy class containing g 2 G is the index of the centraliser, [G : Cg]. Further X jGj = jZ(G)j + [G : Cg]; [G:Cg]>1 where the second sum run over those conjugacy classes with more than one element. Proof. Let G act on itself by conjugation. Then the orbits are the conjugacy classes. If g 2 then the stabiliser of g is nothing more than the centraliser. Thus the cardinality of the conjugacy class containing g is [G : Cg] by (23.3). If g 2 G is in the centre of G then the conjugacy class containing G has only one element, and vice-versa. As G is a disjoint union of its conjugacy classes, we get the second equation. Lemma 23.7. If G is a p-group then the centre of G is a non-trivial subgroup of G. In particular G is simple if and only if the order of G is p. Proof. Consider the class equation X jGj = jZ(G)j + [G : Cg]: [G:Cg]>1 The first and last terms are divisible by p and so the order of the centre of G is divisible by p. In particular the centre is a non-trivial subgroup. If G is not abelian then the centre is a proper normal subgroup and G is not simple. If G is abelian then G is simple if and only if its order is p. Theorem 23.8. Let G be a finite group whose order is divisible by a prime p. Then G contains at least one Sylow p-subgroup. 3 Proof. Suppose that n = pkm, where m is coprime to p. Let S be the set of subsets of G of cardinality pk. Then the cardi- nality of S is given by a binomial n pkm(pkm − 1)(pkm − 2) ::: (pkm − pk + 1) = pk pk(pk − 1) ::: 1 Note that for every term in the numerator that is divisible by a power of p, we can match this term in the denominator which is also divisible by the same power of p. In particular the cardinality of S is coprime to p. Now let G act on S by left translation, G × S −! S where (g; P ) −! gP: Then S is breaks up into orbits. As the cardinality is coprime to p, it follows that there is an orbit whose cardinality is coprime to p. Suppose that X belongs to this orbit. Pick g 2 X and let P = g−1X. Then P contains the identity. Let H be the stabiliser of P . Then H ⊂ P , since h · e 2 P . On the other hand, [G : H] is coprime to p, so that the order of H is divisible by pk. It follows that H = P . But then P is a Sylow p-subgroup. Question 23.9. What is the automorphism group of Sn? Definition-Lemma 23.10. Let G be a group. If a 2 G then conjugation by G is an automorphism σa of G, called an inner automorphism of G. The group G0 of all inner automor- phisms is isomorphic to G=Z, where Z is the centre. G0 is a normal subgroup of Aut(G) the group of all automorphisms and the quotient is called the outer automorphism group of G. Proof. There is a natural map ρ: G −! Aut(G); whose image is G0. The kernel is isomorphic to the centre and so G0 ' G=Z; by the first Isomorphism theorem. It follows that G0 ⊂ Aut(G) is a subgroup. Suppose that φ: G −! G is any automorphism of G.I claim that −1 φσaφ = σφ(a): 4 Since both sides are functions from G to G it suffices to check they do the same thing to any element g 2 G. −1 −1 −1 φσaφ (g) = φ(aφ (g)a ) = φ(a)gφ(a)−1 = σφ(a)(g): 0 Thus G is normal in Aut(G). Lemma 23.11. The centre of Sn is trivial unless n = 2. Proof. Easy check. Theorem 23.12. The outer automorphism group of Sn is trivial unless n = 6 when it is isomorphic to Z2. Lemma 23.13. If φ: Sn −! Sn is an automorphism of Sn which sends a transposition to a transposition then φ is an inner automorphism. Proof. Since any automorphism permutes the conjugacy classes, φ sends transpositions to transpositions. Suppose that φ(1; 2) = (i; j). Let a = (1; i)(2; j). Then σa(i; j) = (1; 2) and so σaφ fixes (1; 2). It is ob- viously enough to show that σaφ is an inner automorphism. Replacing φ by σaφ we may assume φ fixes (1; 2). Now consider τ = φ(2; 3). By assumption τ is a transposition. Since (1; 2) and (2; 3) both move 2, τ must either move 1 or 2. Suppose it moves 1. Let a = (1; 2). Then σaφ still fixes (1; 2) and σaτ moves 2. Replacing φ by σaφ we may assume τ = (2; i), for some i. Let a = (3; i).
Recommended publications
  • Group Homomorphisms
    1-17-2018 Group Homomorphisms Here are the operation tables for two groups of order 4: · 1 a a2 + 0 1 2 1 1 a a2 0 0 1 2 a a a2 1 1 1 2 0 a2 a2 1 a 2 2 0 1 There is an obvious sense in which these two groups are “the same”: You can get the second table from the first by replacing 0 with 1, 1 with a, and 2 with a2. When are two groups the same? You might think of saying that two groups are the same if you can get one group’s table from the other by substitution, as above. However, there are problems with this. In the first place, it might be very difficult to check — imagine having to write down a multiplication table for a group of order 256! In the second place, it’s not clear what a “multiplication table” is if a group is infinite. One way to implement a substitution is to use a function. In a sense, a function is a thing which “substitutes” its output for its input. I’ll define what it means for two groups to be “the same” by using certain kinds of functions between groups. These functions are called group homomorphisms; a special kind of homomorphism, called an isomorphism, will be used to define “sameness” for groups. Definition. Let G and H be groups. A homomorphism from G to H is a function f : G → H such that f(x · y)= f(x) · f(y) forall x,y ∈ G.
    [Show full text]
  • Complete Objects in Categories
    Complete objects in categories James Richard Andrew Gray February 22, 2021 Abstract We introduce the notions of proto-complete, complete, complete˚ and strong-complete objects in pointed categories. We show under mild condi- tions on a pointed exact protomodular category that every proto-complete (respectively complete) object is the product of an abelian proto-complete (respectively complete) object and a strong-complete object. This to- gether with the observation that the trivial group is the only abelian complete group recovers a theorem of Baer classifying complete groups. In addition we generalize several theorems about groups (subgroups) with trivial center (respectively, centralizer), and provide a categorical explana- tion behind why the derivation algebra of a perfect Lie algebra with trivial center and the automorphism group of a non-abelian (characteristically) simple group are strong-complete. 1 Introduction Recall that Carmichael [19] called a group G complete if it has trivial cen- ter and each automorphism is inner. For each group G there is a canonical homomorphism cG from G to AutpGq, the automorphism group of G. This ho- momorphism assigns to each g in G the inner automorphism which sends each x in G to gxg´1. It can be readily seen that a group G is complete if and only if cG is an isomorphism. Baer [1] showed that a group G is complete if and only if every normal monomorphism with domain G is a split monomorphism. We call an object in a pointed category complete if it satisfies this latter condi- arXiv:2102.09834v1 [math.CT] 19 Feb 2021 tion.
    [Show full text]
  • On Automorphisms and Endomorphisms of Projective Varieties
    On automorphisms and endomorphisms of projective varieties Michel Brion Abstract We first show that any connected algebraic group over a perfect field is the neutral component of the automorphism group scheme of some normal pro- jective variety. Then we show that very few connected algebraic semigroups can be realized as endomorphisms of some projective variety X, by describing the structure of all connected subsemigroup schemes of End(X). Key words: automorphism group scheme, endomorphism semigroup scheme MSC classes: 14J50, 14L30, 20M20 1 Introduction and statement of the results By a result of Winkelmann (see [22]), every connected real Lie group G can be realized as the automorphism group of some complex Stein manifold X, which may be chosen complete, and hyperbolic in the sense of Kobayashi. Subsequently, Kan showed in [12] that we may further assume dimC(X) = dimR(G). We shall obtain a somewhat similar result for connected algebraic groups. We first introduce some notation and conventions, and recall general results on automorphism group schemes. Throughout this article, we consider schemes and their morphisms over a fixed field k. Schemes are assumed to be separated; subschemes are locally closed unless mentioned otherwise. By a point of a scheme S, we mean a Michel Brion Institut Fourier, Universit´ede Grenoble B.P. 74, 38402 Saint-Martin d'H`eresCedex, France e-mail: [email protected] 1 2 Michel Brion T -valued point f : T ! S for some scheme T .A variety is a geometrically integral scheme of finite type. We shall use [17] as a general reference for group schemes.
    [Show full text]
  • Orbits of Automorphism Groups of Fields
    ORBITS OF AUTOMORPHISM GROUPS OF FIELDS KIRAN S. KEDLAYA AND BJORN POONEN Abstract. We address several specific aspects of the following general question: can a field K have so many automorphisms that the action of the automorphism group on the elements of K has relatively few orbits? We prove that any field which has only finitely many orbits under its automorphism group is finite. We extend the techniques of that proof to approach a broader conjecture, which asks whether the automorphism group of one field over a subfield can have only finitely many orbits on the complement of the subfield. Finally, we apply similar methods to analyze the field of Mal'cev-Neumann \generalized power series" over a base field; these form near-counterexamples to our conjecture when the base field has characteristic zero, but often fall surprisingly far short in positive characteristic. Can an infinite field K have so many automorphisms that the action of the automorphism group on the elements of K has only finitely many orbits? In Section 1, we prove that the answer is \no" (Theorem 1.1), even though the corresponding answer for division rings is \yes" (see Remark 1.2). Our proof constructs a \trace map" from the given field to a finite field, and exploits the peculiar combination of additive and multiplicative properties of this map. Section 2 attempts to prove a relative version of Theorem 1.1, by considering, for a non- trivial extension of fields k ⊂ K, the action of Aut(K=k) on K. In this situation each element of k forms an orbit, so we study only the orbits of Aut(K=k) on K − k.
    [Show full text]
  • The General Linear Group
    18.704 Gabe Cunningham 2/18/05 [email protected] The General Linear Group Definition: Let F be a field. Then the general linear group GLn(F ) is the group of invert- ible n × n matrices with entries in F under matrix multiplication. It is easy to see that GLn(F ) is, in fact, a group: matrix multiplication is associative; the identity element is In, the n × n matrix with 1’s along the main diagonal and 0’s everywhere else; and the matrices are invertible by choice. It’s not immediately clear whether GLn(F ) has infinitely many elements when F does. However, such is the case. Let a ∈ F , a 6= 0. −1 Then a · In is an invertible n × n matrix with inverse a · In. In fact, the set of all such × matrices forms a subgroup of GLn(F ) that is isomorphic to F = F \{0}. It is clear that if F is a finite field, then GLn(F ) has only finitely many elements. An interesting question to ask is how many elements it has. Before addressing that question fully, let’s look at some examples. ∼ × Example 1: Let n = 1. Then GLn(Fq) = Fq , which has q − 1 elements. a b Example 2: Let n = 2; let M = ( c d ). Then for M to be invertible, it is necessary and sufficient that ad 6= bc. If a, b, c, and d are all nonzero, then we can fix a, b, and c arbitrarily, and d can be anything but a−1bc. This gives us (q − 1)3(q − 2) matrices.
    [Show full text]
  • Arxiv:Math/0511624V1
    Automorphism groups of polycyclic-by-finite groups and arithmetic groups Oliver Baues ∗ Fritz Grunewald † Mathematisches Institut II Mathematisches Institut Universit¨at Karlsruhe Heinrich-Heine-Universit¨at D-76128 Karlsruhe D-40225 D¨usseldorf May 18, 2005 Abstract We show that the outer automorphism group of a polycyclic-by-finite group is an arithmetic group. This result follows from a detailed structural analysis of the automorphism groups of such groups. We use an extended version of the theory of the algebraic hull functor initiated by Mostow. We thus make applicable refined methods from the theory of algebraic and arithmetic groups. We also construct examples of polycyclic-by-finite groups which have an automorphism group which does not contain an arithmetic group of finite index. Finally we discuss applications of our results to the groups of homotopy self-equivalences of K(Γ, 1)-spaces and obtain an extension of arithmeticity results of Sullivan in rational homo- topy theory. 2000 Mathematics Subject Classification: Primary 20F28, 20G30; Secondary 11F06, 14L27, 20F16, 20F34, 22E40, 55P10 arXiv:math/0511624v1 [math.GR] 25 Nov 2005 ∗e-mail: [email protected] †e-mail: [email protected] 1 Contents 1 Introduction 4 1.1 Themainresults ........................... 4 1.2 Outlineoftheproofsandmoreresults . 6 1.3 Cohomologicalrepresentations . 8 1.4 Applications to the groups of homotopy self-equivalences of spaces 9 2 Prerequisites on linear algebraic groups and arithmetic groups 12 2.1 Thegeneraltheory .......................... 12 2.2 Algebraicgroupsofautomorphisms. 15 3 The group of automorphisms of a solvable-by-finite linear algebraic group 17 3.1 The algebraic structure of Auta(H)................
    [Show full text]
  • MATH 412 PROBLEM SET 11, SOLUTIONS Reading: • Hungerford
    MATH 412 PROBLEM SET 11, SOLUTIONS Reading: • Hungerford 8.4, 9.1, 9.2 Practice Problems: These are “easier” problems critical for your understanding, but not to be turned in. Be sure you can do these! You may see them or something very similar on Quizzes and Exams. Section 8.4: 9, 19, 21, Section 9.1: 1, 2, 3, 4, 5 Section 9.2: 1, 3 Problems due Friday. Write out carefully, staple solutions to A, B and C in one pack, and stable solutions to D and E in another. Write your name and section number on each. (A) Let F be a field. Consider the group SL2(F) of 2 × 2 matrices with entries in F and determinant one. 1 (1) Show that the center of SL2(F) is Z = {±Ig, where I is the 2 × 2 identity matrix. (2) Let PSL2(F) be the quotient group SL2(F)=Z. Compute the order of PSL2(Z7) as a group. 2 (3) For each of the following, compute its order as an element of PSL2(Z7). 0 2 • A = ± 3 0 1 1 • B = ± 6 0 1 3 • C = ± 0 1 Note: PSL2(Z7) is the smallest simple group that is neither cyclic nor isomorphic to An for some n. Solution. a b 1 1 1 0 (1) Let A = be an arbitrary element of SL ( ), B = , and C = . We c d 2 F 0 1 1 1 −c a − d b 0 compute AB − BA = and AC − CA = . If A 2 Z, then AB − BA = 0 c d − a −b AC − CA = 0, so b = c = a − d = 0.
    [Show full text]
  • The Pure Symmetric Automorphisms of a Free Group Form a Duality Group
    THE PURE SYMMETRIC AUTOMORPHISMS OF A FREE GROUP FORM A DUALITY GROUP NOEL BRADY, JON MCCAMMOND, JOHN MEIER, AND ANDY MILLER Abstract. The pure symmetric automorphism group of a finitely generated free group consists of those automorphisms which send each standard generator to a conjugate of itself. We prove that these groups are duality groups. 1. Introduction Let Fn be a finite rank free group with fixed free basis X = fx1; : : : ; xng. The symmetric automorphism group of Fn, hereafter denoted Σn, consists of those au- tomorphisms that send each xi 2 X to a conjugate of some xj 2 X. The pure symmetric automorphism group, denoted PΣn, is the index n! subgroup of Σn of symmetric automorphisms that send each xi 2 X to a conjugate of itself. The quotient of PΣn by the inner automorphisms of Fn will be denoted OPΣn. In this note we prove: Theorem 1.1. The group OPΣn is a duality group of dimension n − 2. Corollary 1.2. The group PΣn is a duality group of dimension n − 1, hence Σn is a virtual duality group of dimension n − 1. (In fact we establish slightly more: the dualizing module in both cases is -free.) Corollary 1.2 follows immediately from Theorem 1.1 since Fn is a 1-dimensional duality group, there is a short exact sequence 1 ! Fn ! PΣn ! OPΣn ! 1 and any duality-by-duality group is a duality group whose dimension is the sum of the dimensions of its constituents (see Theorem 9.10 in [2]). That the virtual cohomological dimension of Σn is n − 1 was previously estab- lished by Collins in [9].
    [Show full text]
  • GROUP ACTIONS 1. Introduction the Groups Sn, An, and (For N ≥ 3)
    GROUP ACTIONS KEITH CONRAD 1. Introduction The groups Sn, An, and (for n ≥ 3) Dn behave, by their definitions, as permutations on certain sets. The groups Sn and An both permute the set f1; 2; : : : ; ng and Dn can be considered as a group of permutations of a regular n-gon, or even just of its n vertices, since rigid motions of the vertices determine where the rest of the n-gon goes. If we label the vertices of the n-gon in a definite manner by the numbers from 1 to n then we can view Dn as a subgroup of Sn. For instance, the labeling of the square below lets us regard the 90 degree counterclockwise rotation r in D4 as (1234) and the reflection s across the horizontal line bisecting the square as (24). The rest of the elements of D4, as permutations of the vertices, are in the table below the square. 2 3 1 4 1 r r2 r3 s rs r2s r3s (1) (1234) (13)(24) (1432) (24) (12)(34) (13) (14)(23) If we label the vertices in a different way (e.g., swap the labels 1 and 2), we turn the elements of D4 into a different subgroup of S4. More abstractly, if we are given a set X (not necessarily the set of vertices of a square), then the set Sym(X) of all permutations of X is a group under composition, and the subgroup Alt(X) of even permutations of X is a group under composition. If we list the elements of X in a definite order, say as X = fx1; : : : ; xng, then we can think about Sym(X) as Sn and Alt(X) as An, but a listing in a different order leads to different identifications 1 of Sym(X) with Sn and Alt(X) with An.
    [Show full text]
  • Generalized Quaternions
    GENERALIZED QUATERNIONS KEITH CONRAD 1. introduction The quaternion group Q8 is one of the two non-abelian groups of size 8 (up to isomor- phism). The other one, D4, can be constructed as a semi-direct product: ∼ ∼ × ∼ D4 = Aff(Z=(4)) = Z=(4) o (Z=(4)) = Z=(4) o Z=(2); where the elements of Z=(2) act on Z=(4) as the identity and negation. While Q8 is not a semi-direct product, it can be constructed as the quotient group of a semi-direct product. We will see how this is done in Section2 and then jazz up the construction in Section3 to make an infinite family of similar groups with Q8 as the simplest member. In Section4 we will compare this family with the dihedral groups and see how it fits into a bigger picture. 2. The quaternion group from a semi-direct product The group Q8 is built out of its subgroups hii and hji with the overlapping condition i2 = j2 = −1 and the conjugacy relation jij−1 = −i = i−1. More generally, for odd a we have jaij−a = −i = i−1, while for even a we have jaij−a = i. We can combine these into the single formula a (2.1) jaij−a = i(−1) for all a 2 Z. These relations suggest the following way to construct the group Q8. Theorem 2.1. Let H = Z=(4) o Z=(4), where (a; b)(c; d) = (a + (−1)bc; b + d); ∼ The element (2; 2) in H has order 2, lies in the center, and H=h(2; 2)i = Q8.
    [Show full text]
  • On the Group-Theoretic Properties of the Automorphism Groups of Various Graphs
    ON THE GROUP-THEORETIC PROPERTIES OF THE AUTOMORPHISM GROUPS OF VARIOUS GRAPHS CHARLES HOMANS Abstract. In this paper we provide an introduction to the properties of one important connection between the theories of groups and graphs, that of the group formed by the automorphisms of a given graph. We provide examples of important results in graph theory that can be understood through group theory and vice versa, and conclude with a treatment of Frucht's theorem. Contents 1. Introduction 1 2. Fundamental Definitions, Concepts, and Theorems 2 3. Example 1: The Orbit-Stabilizer Theorem and its Application to Graph Automorphisms 4 4. Example 2: On the Automorphism Groups of the Platonic Solid Skeleton Graphs 4 5. Example 3: A Tight Bound on the Product of the Chromatic Number and Independence Number of Vertex-Transitive Graphs 6 6. Frucht's Theorem 7 7. Acknowledgements 9 8. References 9 1. Introduction Groups and graphs are two highly important kinds of structures studied in math- ematics. Interestingly, the theory of groups and the theory of graphs are deeply connected. In this paper, we examine one particular such connection: that which emerges from the observation that the automorphisms of any given graph form a group under composition. In section 2, we provide a framework for understanding the material discussed in the paper. In sections 3, 4, and 5, we demonstrate how important results in group theory illuminate some properties of automorphism groups, how the geo- metric properties of particular embeddings of graphs can be used to determine the structure of the automorphism groups of all embeddings of those graphs, and how the automorphism group can be used to determine fundamental truths about the structure of the graph.
    [Show full text]
  • 14.2 Covers of Symmetric and Alternating Groups
    Remark 206 In terms of Galois cohomology, there an exact sequence of alge- braic groups (over an algebrically closed field) 1 → GL1 → ΓV → OV → 1 We do not necessarily get an exact sequence when taking values in some subfield. If 1 → A → B → C → 1 is exact, 1 → A(K) → B(K) → C(K) is exact, but the map on the right need not be surjective. Instead what we get is 1 → H0(Gal(K¯ /K), A) → H0(Gal(K¯ /K), B) → H0(Gal(K¯ /K), C) → → H1(Gal(K¯ /K), A) → ··· 1 1 It turns out that H (Gal(K¯ /K), GL1) = 1. However, H (Gal(K¯ /K), ±1) = K×/(K×)2. So from 1 → GL1 → ΓV → OV → 1 we get × 1 1 → K → ΓV (K) → OV (K) → 1 = H (Gal(K¯ /K), GL1) However, taking 1 → µ2 → SpinV → SOV → 1 we get N × × 2 1 ¯ 1 → ±1 → SpinV (K) → SOV (K) −→ K /(K ) = H (K/K, µ2) so the non-surjectivity of N is some kind of higher Galois cohomology. Warning 207 SpinV → SOV is onto as a map of ALGEBRAIC GROUPS, but SpinV (K) → SOV (K) need NOT be onto. Example 208 Take O3(R) =∼ SO3(R)×{±1} as 3 is odd (in general O2n+1(R) =∼ SO2n+1(R) × {±1}). So we have a sequence 1 → ±1 → Spin3(R) → SO3(R) → 1. 0 × Notice that Spin3(R) ⊆ C3 (R) =∼ H, so Spin3(R) ⊆ H , and in fact we saw that it is S3. 14.2 Covers of symmetric and alternating groups The symmetric group on n letter can be embedded in the obvious way in On(R) as permutations of coordinates.
    [Show full text]